Vous êtes sur la page 1sur 34

NASA Technical Memorandum 1043_3 ?_ ......

Calculation of a Circular Jet in Crossflow


With a Multiple-Time-Scale
Turbulence Model
S.-W. Kim
University of Texas at Arlington ....
Arlington, Texas .......
and
T.J. Benson
Lewis Research Center
Cleveland, Ohio
July 1991 ..............................................
7
n
(NA_A-TM-104345) CALCULATI..,N 0r A CIRCULAR
JFI IN CRqSSFLUW WITH A MULTIPLE-TIME-SCALE
TU_nUL_NCF MGD_L (NASA) 32 p CSCL 20D
L..
G3/34
"r _.,2 L__;
N91-30670
Uncl as
O0 3 7 _ _ 3
CALCULATION OF A CIRCULAR JET IN CROSSFLOW WITH A
MULTIPLE-TII_-SCALE TURBULANCE MODEL
S.-W. Kim
University of Texas at Arlington
Department of Aerospace Engineering
Arlington, Texas 76010
and
T.J. Benson
National Aeronautics and Space Administration
Lewis Research Center
Cleveland, Ohio 44135
SUMMARY
Numerical calculation of a three-dlmenslonal turbulent flow of a Jet in a
crossflow using a multiple-time-scale turbulence model is presented. The
turbulence in the forward region of the Jet is in a stronger inequilibrium
state than that in the wake region of the Jet, while the turbulence level
in the wake region is higher than that in the front region. The calculated
flow and the concentration fields are in very good agreement with the
measured data, and it indicates that the turbulent transport of mass,
concentration and momentum is strongly governed by the inequilibrium
turbulence. The capability of the multiple-tlme-scale turbulence model to
resolve the inequilibrium turbulence field is also discussed.
*NASA Resident Research Associate at Lewis Research Center.
Aj
c
Cpl
ctl
c_
c_f
D
i,j
k
kp
kt
P
p'
Pr
Uo
uj
wj
y+
Cp
Et
_d
_t
P
a d
al
., ........ NOMENCLATURE
coefficientfor uj-velocity correction
normalized concentration
turbulence model constants for _p equation (2-1,3)
turbulence model constants for _t equation (2-1,3)
eddy viscosity coefficient
constant coefficient (-0.09)
diameter of the circular Jet
index for spatial coordinate (i-i,2,3 and j-1,2,3)
turbulent kinetic energy (k-_+k t)
turbulent kinetic energy in production range
turbulent kinetic energy in dissipation range
pressure
incremental pressure
production rate
free-stream velocity of crossflow
time averaged velocity (-{u,v,w})
jet velocity averaged across jet cross-section
wall coordinate based on friction velocity
energy transfer rate
dissipation rat6
molecular viscosity
molecular diffusivity
turbulent viscosity
density
turbulent Schmidt number
turbulent Prandtl number for 2-equatlon, 2-{kp,_p,kt,_ t}
2
INTRODUCTION
Numerical calculation of a circular Jet exhausting into a cross flow [I]
is presented. The circular jet in a uniform crossflow is schematically
shown in Fig. i. Turbulent flows similar to a jet in a crossflow can be
found in a number of engineering applications. For example, in gas turbine
combustors, a number of circumferentially distributed jets are used to
ensure correct combustion in the flame zone and then to dilute the hot gas
entering the turbine. Experimental investigations of air jets in crossflows
have been made to better understand turbulent flows in such engineering
applications even though air Jets in crossflows are by far simpler than
those in a gas turbine combustor. Compilations of various experimental
investigations of jets in crossflows can be found in Crabb et al. [I] and
Khan [2].
With the recent advances in numerical methods to solve the Navier-Stokes
equations, a number of numerical simulations of jets in crossflows have
been reported. A compilation of various numerical investigations of the
flows can be found in Claus and Vanka [3]. Earlier numerical calculations
of the flows [4,5] have been devoted to the development and the
verification of the numerical methods, and only a very small number of grid
points were used to discretize the entire flow domain due to the limited
capability of computers. More recently, a realistic number of grid points
began to be used to calculate the flows using k-c turbulence models [6].
The numerical results obtained using the relatively fine meshes show
improved comparison with the measured data in a certain part of the flow
domain and worse agreement with the measured data in the other part of the
flow domain. Claus and Vanka [3] carried out a grid independence study of a
jet in a crossflow using a k-_ turbulence model to identify the cause of
the deteriorated comparison with the measureddata. Claus and Vanka [3]
showedthat the numerical results obtained using a 128x48x48meshare not
significantly different from those obtained using a 256x96x96meshand that
the deteriorated comparison is caused by the inability of the k-_
turbulence model to describe the complex turbulence field. In each of the
above numerical simulations, the upstream region of the Jet was excluded
from the _omputatlonal domain. However, Andreopoulos [7] showed that the
circular jet and the crossflow interact s@rongly with each other at the jet
exit and that the influence is propagated toward the upstream region of the
jet. Thus, the deteriorated numerical results can also be caused by the
numerical models which can not fully account for the strong interaction
between the jet and the crossflow. In the present numerical study, the
boundary for the circular Jet is located at one diameter upstream of the
jet exit so that the strong interaction at the Jet exit is also accurately
simulated.
Numerical results for various complex turbulent flows obtained using
two-equatlon turbulence models, algebraic Reynolds stress turbulence models
(ARSM) and Reynolds stress turbulence models (RSM) show that these
turbulence models can not accurately describe the turbulence fields of
various complex turbulent flows [8]. One common inability of the
two-equatlon turbulence models, ARSM and RSM is that these turbulence
models can not account for "inequillbrium turbulence" due to the use of a
single time scale to describe both the turbulent transport and the
dissipation of the turbulent kinetic energy. The "inequillbrlum turbulence"
is explained in the "Multlple-Time-scale Turbulence Model" section. The
multiple-time-Scale turbulence model [9] (hereafter abbreviated as the M-S
turbulence model for convenience) yields accurate numerical results for
widely different classes of complex turbulent flows (e.g., turbulent flows
subjected to extra strains causedby streamline curvatures, interaction of
multiple numberof turbulence fields, and shock wave-turbulent boundary
layer interactions). The complex turbulent flows to which the present M-S
turbulence model has been applied as yet include a wall-jet flow, a shear
layer with wake-boundary layer interaction, a backward-facing step flow, a
confined coaxial swirling jet, turbulent shear layers over curved surfaces,
separated transonic turbulent flows over a curved hill and reattaching
shear layers in a divergent channel. It can be seen in [9-12] that the
numerical results for these complex turbulent flows obtained using the M-S
turbulence model are in as good agreement with the measured data as those
obtained using an optimized k-_, ARSM, or RSM turbulence model for each
flow case. The capability of the M-S turbulence model to solve widely
different complex turbulent flows is attributed to its capability to
resolve the inequilibrium turbulence. This capability is discussed in the
"Multiple-time-Scale Turbulence Model" section.
The fluid flow in the near-wall region of the jet exit is subjected to a
large mean flow strain rate. The near-wall turbulence field, intensified by
the large mean flow strain rate, can influence the entire fluid flow in the
downstream region of the jet. Thus the near-wall turbulence field in the
vicinity of the jet exit needs to be resolved accurately in order to
correctly predict the entire flow field. In the present numerical
simulation, the near-wall turbulence is described by a "partially low
Reynolds number" near-wall turbulence model [13]. In the model, only the
turbulent kinetic energy equations are extended to include the near-wall
low turbulence region and the energy transfer rate and the dissipation rate
inside the near-wall layer are obtained from algebraic equations. It was
shownin numerical calculations of turbulent flows over curved hills [I0],
transonic turbulent flows over a curved surface with shock wave-boundary
layer interactions [II], and reattaching shear layers in a divergent
channel [12] that the "partially low-Reynolds number" near-wall turbulence
model (when used together with the M-S turbulence model) accurately
predicts the near-wall turbulence fields.
The present numerical method is a finite volume method based on a
pressure correction algorithm [14-16]. In the method, all flow variables,
except pressure and concentration, are located at the same grid points,
while pressure and concentration are located at the centroid of a cell
formed by the eight neighboring velocity grid points. The pressure
correction algorithm is described in the following section. Calculations of
a three-dimenslonal lld-drlven cavity flow and a laminar flow through a
90-bend square duct can be found in [15]. It is shown in [15] that the
numerical results for the cavity flow obtained using the present numerical
method compare more favorably with the measured data than those obtained
using a formally third order accurate quadratic upwind interpolation
scheme. It is also shown in [15] that the present method yields a grid
independent solution for the curved duct flow with a very small number of
grid points and that the method yields quickly and strongly convergent
numerical results. Application of the same numerical method for
two-dlmenslonal flows can be found in [i0-12].
NUMERICAL METHOD
The incompressible turbulent flow equations are given as;
a
--(puj) - o. (I)
axj
a a [ au i auj)l "
ax-_.(#uiuj) - a_jt(P+"t)(_ + ax'-'_ j
a 2
---(p + -pk)
ax i 3
(2)
where repeated indices imply summation over the index unless otherwise
stated. The convectlon-diffuslon equation for the concentration is given
as,
a a[ .tac]
a_j(pujc) -I_(#d+i)-- _ - 0 (3)
axj[ ad axjj
where ad=0.75 is used in the present study. Due to the strong large eddy
mixing, the molecular diffusivlty can be ignored or formally approximated
as #d-#/ad; and neither of the approximations influence the numerical
results significantly.
In the numerical method, the conservation of mass equation is replaced by
a pressure correction equation given as:
__a [_j ap,] _ ap*uj* (4)
axj q) axj
where uj* denotes the velocity which may not satisfy the conservation of
mass as yet and the last term represents the mass imbalance (see [15] for
details). The flow equations, the concentration equation, and the
turbulence equations are solved by a finite volume method using the
power-law upwind differencing scheme [14]. As all the central-dlfferenced
finite volume equations for self-adjolnt second order elliptic partial
differential equations are strongly diagonally dominant, the discrete
pressure correction equation obtained by applying the standard finite
volume method to eq. (4) is strongly diagonally dominant even for highly
skewed and graded meshes.
For completeness, the veloclty-pressure decoupling that occurs when
various pressure correction algorithms are used for pressure-staggered
meshes is described briefly below. The use of various pressure correction
algorithms in pressure-staggered meshes does not yield a diagonally
dominant system of equations for the incremental pressure. In such a case,
the mass imbalance at a pressure grid point produces large corrections for
pressures at adjacent pressure grid points and veloclty-pressure decoupling
occurs [15-16]. For clarity, the differences between the present pressure
correction algorithm and various other algorithms are summarized below. For
an orthogonal mesh aligned with cartesian coordinates, the present pressure
correction algorithm and the "momentum interpolation scheme" of Perlc et
al. [17] and Majumdar [18] yield a 7-diagonal system of equations for the
incremental pressure, while various other pressure correction algorithms
yield a 27-diagonal system of equations which lacks diagonal dominance
[15-16]. In the momentum interpolation scheme, the pressure gradient is
interpolated differently from the other terms in the discrete momentum
equation to achieve diagonal dominance. In curvilinear coordinates, the
present pressure correction algorithm yields a 27-diagonal system of
equations, while the momentum interpolation scheme always yields a
7-diagonal system of equations and it can not account for grid skewness in
the discrete pressure correction equation in its present form. Also a
specialized interpolation scheme needs to be adopted to obtain a unique
solution that does not depend on the under-relaxatlon parameter [18]. On
the other hand, the present pressure correction method yields a unique
solution_slnce the incremental pressure is driven only by the mass
imbalance as shown in eq. (4).
8
MULTIPLE-TIME-SCALE TURBULENCE MODEL
The anlsotropy of the turbulence is the most easily detectable phenomenon
in a measurement of a turbulent flow. Thus, it was conceived that the poor
capability of the two-equation turbulence models to resolve complex
turbulent flows is attributed to the inability of the turbulence models to
take into account of the anlsotropy of the turbulence. Thus the emphasis is
lald upon improving the ARSM and the RSM. However, a number of numerical
investigations carried out during the last one and half decades show that
the ARSM and RSM still can not accurately predict the turbulence phenomena
occurring in various flows unless the pressure-straln rate correlation is
optimized for each flow [11-12].
A careful examination of semi-emplrlcal data (theoretically derived data
from a set of measured data) reveals that the inequilibrium turbulence also
dictates the developments of the mean flow field and the turbulence field.
Here, the "inequilibrlum turbulence" represents the state of a turbulence
field in which Pr/_ t varies rapidly in space so that the shape and the
frequency domain of the spectral density varies widely in space. The
spectral density curves shown in Fig. 2-(a) are constructed based on the
measured data of Klebanoff [19] and Wygnanskl and Fiedler [20]. It can also
be seen in Fig. 2-(a) that the generation of the energy containing large
eddies by the instability of the mean fluid motion occurs in the low
frequency region and that the peak of the spectral density moves toward the
high frequency region as Pr/_ t is decreased. The seml-emplrlcal c_ for a
plane jet obtained by Rodl [21] is shown in Fig. 3. It can be seen in the
figure that c_ is decreased as Pr/_ t is increased, and c_ is increased as
Pr/_ t is decreased. Thus, the developments of the mean fluid flow and the
9
turbulence field are influenced by the spatially varying turbulent eddy
viscosity, and the spatially varying turbulent eddy viscosity depends on
the level of the local inequillbrlum turbulence. Consider k-_ turbulence
models for which the eddy viscosity is given as _t-pc#fk2/_ t. Due to the
use of a constant c_f, the eddy viscosity in a strongly turbulent region
(Pr/_t>l) is over-predlcted and that in weakly turbulent region (Pr=O) is
under-predicted. As an example, the under-predicted reattachment location
of a backward-facing step flow obtained using a k-_ turbulence model is
caused by the over-predicted turbulent viscosity along the reattaching
shear layer [12].
The variation of c_ as a function of Pr/_ t was incorporated into k-_
turbulence models in the form of "generalized algebraic stress turbulence
models." The c_ curves by Launder [22] and Kim and Chen [23] are shown in
Fig. 3. The generalized algebraic stress turbulence models yield accurate
numerical results for shear layers when used in boundary layer flow
solvers. However, the use of these turbulence models in elliptic
(two-dimensional) flow solvers does not easily yield a converged solution
due to a severe interpolation used in the c9 function. Furthermore, it is
not clear if the generalized algebraic stress turbulence models can resolve
the inequilibrlum turbulence as cleanly as the M-S turbulence models can,
since the generalized algebraic stress turbulence models lack many features
of the M-S turbulence models to be described below.
The M-S turbulence models [9,24] appeared as a consequence of recognizing
the inability of various slngle-time-scale turbulence models (k-_, ARSM,
and RSM) to accurately describe complex turbulent flows. The
convectlon-diffusion equations of the M-S turbulence models were
established by Hanjellc et al. [24]. The convectlon-diffuslon equations
I0
most naturally describe the physically observed turbulence phenomenain the
sense that the turbulent transport of mass and momentumis described using
the time-scale of large eddies and the dissipation rate is described using
the time-scale of fine-scale eddies. Later, Kim and Chen [9] established
the general form of the load functions based on a physical dimensional
analysis. The differences between the present M-S turbulence model and that
of Hanjelic et al. [24] can be found in [11-12], and hence, these are not
repeated here. The capability of the M-Sturbulence model to solve complex
turbulent flows was further enhanced by incorporating a "partially
low-Reynolds number" near-wall turbulence model into the M-S turbulence
model [10-13]. Calculations of widely different classes of complex
turbulent flows showed that the M-S turbulence model can resolve the
inequillbrlum turbulence and can model the cascade of turbulent kinetic
energy. These capabilities of the M-S turbulence model are described below.
The M-S turbulence equations are given below for completeness. The
turbulent kinetic energy and the energy transfer rate equations for energy
containing large eddies are given as;
+ - pPr - p_p
puj axj axj [ akp
aEp a__ _t a_pl pr2 Pr(p (p2
PU3axj axj[ (_ + a_)_p axjj--- pCplI_ + pCp2_kp - pcp3kp
where the production rate is given as;
(5)
(6)
Pr- 2 2 -- 2 + + -- + +
[ayj [azj axj
The turbulent kinetic energy and the dissipation rate equations for fine
ii
scale eddies are given as:
O { aktl-
-(,+"t ) _xjj
axj akt
Oe t
pUJaxj
a { _t a_t 1
-- (_+--) -- I
axj a_t axjj
Pep - P_t
2
Cp CpCt
PCtl_ + PCt2_ -
kt k t
2
_t
PCt3
kt
(7)
(8)
and the eddy viscosity is given as;
k2
_t I pc_f--
_p
(9)
The turbulence model constants are given as; akp-0.75 , akt-0.75 , acp-l.15,
actIl.15, Cpl-0.21, Cp2-1.24, Cp3- 1.84, Ctli0.29, ct21 1.28, and ct3-1.66.
The capability of the M-S turbulence model to resolve the Inequilibrium
turbulence depends largely on the load functions of the _p and _t equations
and the way the turbulence model constants are established. The load
functions of the Cp and Et equations are obtained from a physical
dimensional analysis [9], and the establishment of the model constants are
based on the assumptions that the turbulence field of a uniformly sheared
flow can approach an asymptotic state in which Pr/_ t becomes a constant and
that the ratio of ct/_p depends on the ratio of Pr/c t. The first assumption
that such an asymptotic state can exist was shown by Harris et al. [25],
and later, was confirmed by Tavoularis and Karnlk [26]. In such asymptotic
states of uniformly sheared flows, the diffusion term vanishes, and the
asymptotic ratio of kp/k t can be obtained by dividing eq. (5) by eq. (7),
i.e,!
kp Dkp Pr/t - Cp/t
kt Dk t _p/E t - i
(i0)
12
It can be seen in eq. (I0) that the existence of the asymptotic ratio of
_/k t depends on the realizability of the second assumption. The second
assumption that the ratio of _t/_p depends on the ratio of Pr/_ t can be
verified by numerical results posterily, or it can be verified indirectly
by comparing the M-S eddy viscosity equation with that of the generalized
algebraic stress turbulence models shown in Fig. 3. The eddy viscosity, eq.
(9), can be rewritten in a form compatible with that of the generalized
algebraic stress turbulence models, i.e.,
k2
_t - pc_--
_t
(Ii)
where c_-c_f(_t/_p) is the eddy viscosity coefficient, the variation of c#
is described by the ratio of _t/_p, and _t/_p is a function of Pr/_t. Thus
the second assumption can be Justified within the context of the
generalized algebraic stress turbulence models.
The three inequilibrium turbulence levels (A, B, and C) imbedded into the
M-S turbulence model (or used in determining the turbulence model
constants) are also shown in Fig. 3. The measured data that corresponds to
the point A in Fig. 3 (i.e., Pr/_t-l.5 ) can be found in Harris et al. [25]
and in Tavoularis and Karnik [26]. The value of _t/_p-0.95 can be estimated
from Fig. 3. The ratio of _/k t for the data point A is obtained to be 9.0
from eq. (I0). For turbulent flows in an equilibrium state (point B in Fig.
3), Pr=_t, and _p has to be equal to both of them to maintain the
equilibrium state. In this case, eq. (i0) becomes indeterminate; and the
ratio of _/kt=4.0 can be obtained from a near-wall analysis of turbulent
flows in equilibrium state [9]. In the free stream region of turbulent
13
flows, the production rate vanishes. Such a case is represented by Point C
in Fig. 3 and the ratio of _t/_p=2.5 can be estimated from the generalized
algebraic stress turbulence model of Kim and Chen [23]. The ratio of
kp/kt=0.7 can be obtained from eq (i0).
The three ratios of _/k t obtained in the above analysis show that
(12)
where A, B and C denote each turbulence state marked in Fig. 3. The
implication of the above inequality is illustrated in Fig. 2-(b), where the
frequency domain is divided into two parts by the simplified spilt-spectrum
[9,24]. The spectral density A in Fig. 2-(b) belongs to a production
dominated region (Pr/_t>l), B belongs to a near equilibrium region
(Pr/_t=l), and C belongs to a region where production rate vanishes (i.e.,
free stream region). The spectral density curves A, B, and C also describe
the cascade process of turbulent kinetic energy that the energy containing
large eddies (characterlzed by low frequencies) are generated by the
instability of the m_an fluid flow, the large eddies become finer eddies
(characterized by higher frequencies), and the fine scale eddies are
dissipated by the viscous force. It can be seen in Fig. 2-(b) that kp/k t is
greater for larger eddies and _/k t becomes smaller for finer-scale eddies.
Thus the ratio of _/k t is determined by the shape and the frequency domain
of each spectral density.
Calculations of various shear layers (boundary layer flows) using the M-S
turbulence model always reproduce the imbedded inequilibrium turbulence
states. For highly complex turbulent flows, large eddies generated in the
upstream region are convected in the downstream directlon_ In this
14
downstream region, the relationship between Pr/_t and _t/_p is influenced
by the convected eddies (i.e., the large value of kp/kt) and the numerical
results exhibit the trend of imbedded inequilibrlum turbulence.
NUMERICAL RESULTS
The flow domain of the jet in a uniform crossflow [I] is shown in Fig. I.
The jet velocity averaged across the cross-sectlonal area of the pipe is
27.6 m/sec, the free stream velocity of the cross flow is 12 m/sec, and the
diameter of the jet is 0.0254 m. In the experiment [i], the concentration
field was measured by injecting helium gas (He) into the circular Jet. The
concentration of the helium is one percent of the alr-hellum mixture at the
Jet exit, and hence the concentration equation, eq. (3), is solved
uncoupled from the momentum equations. The symmetric half of the flow
domain is discretized by 148x61x94 grid points in x-, y-, and z-coordinate
directions, respectively. The body-fitted grid near the jet exit is shown
in Fig. 4. The smallest mesh size in the direction normal to the wall is
0.6x10 "4 m (y+=l.5 based on the fully developed pipe flow) and this mesh
size is sufficiently small to resolve the near-wall turbulence field in the
vicinity of the jet exit. The largest mesh size used near the far field
boundaries is approximately half of the jet diameter.
The inlet boundary conditions for the tangential velocity, the turbulent
kinetic energies, and the dissipation rates (_p and _t) are obtained from
measured data for a fully developed boundary layer flow over a flat plate
[19]. The non-dlmenslonal velocity and the turbulent kinetic energy
profiles are scaled to yield a boundary layer thickness of 0.005 m at the
inlet boundary. The no-sllp boundary condition for velocities, vanishing
gradient for concentration, and vanishing turbulent kinetic energy are
15
prescribed at the solid wall boundary. A vanishing gradient boundary
condition is used for all flow variables at the exit boundary. The
symmetric boundary condition is used on the south (y-O) boundary, and the
free stream boundary condition is used on the north (y-3.1D) and the top
(z-7.bD) boundaries.
The bottom boundary is located at one Jet diameter upstream of the jet
exit (z--ID), and a fully developed pipe flow and a constant concentration
(c-l.O) conditions are prescribed as the Jet inlet boundary. These boundary
conditions can greatly reduce the uncertainty that can be caused by
ill-posed boundary conditions at the Jet inlet as discussed in the
following paragraph. The use of the free stream boundary conditions on the
north and the top boundaries is not a good approximation of the actual
fluid flow unless these boundaries are located sufficiently far away from
the Jet exit. However, it can be seen in the following that the numerical
results near the jet exit are not influenced too much by the far field
boundary conditions [I0]. The partition between the near-wall layer and the
external region is located at y+=100 (based on the fully developed flat
plate flow) and i0 grid points are allocated inside the near-wall layer.
The converged solutions are obtained in approximately 1200 iterations, and
the relative mass and concentration imbalances are 2.5xi0 "5 and 9.5xi0 "3,
respectively.
The contour plots of the jet velocity, the pressure, and the total
pressure at the jet exit are shown in Fig. 5, where the increments between
the contour lines are the same for each contour plot. It can be seen in the
figure that the jet velocity, the static pressure, and the total pressure
vary widely across the cross-sectlon. It does not seem possible to
prescribe a correct boundary condition for the jet if the bottom boundary
16
is located at the Jet exit. For example, in previous numerical calculations
of jets in crossflows, either a constant vertical velocity or a constant
total pressure was prestribed at the jet exit [2,&]. However, the present
numerical results show that a significant amount of uncertainty can be
caused by the use of either of these boundary conditions at the Jet exit.
The calculated velocity vectors, pressure, turbulent kinetic energy,
Pr/_t, _t/ep and _/k t are shown in Fig. 6. The velocity vector and the
pressure contour plot show that the crossflow is decelerated rapidly by the
jet and thus the pressure is increased in the forward region of the jet.
Otherwise, the velocity vector and the pressure contour plot do not show
that any significant phenomena occur in the forward region. However, the
complex Pr/e t and _t/_p contours show that the turbulence field is
experiencing an enormous evolution in the forward region. These contour
plots show that the inequilibrlum turbulence becomes stronger as a fluid
particle approaches the jet and that the peak inequillbrium state occurs
along the interface of the jet and the crossflow. The urbulent kinetic
energy, the production rate, the energy transfer rate and the dissipation
rate in the wake region of the jet are by far greater than those in the
forward region of the jet. However, the turbulence in the forward region is
in a stronger inequilibrlum state than that in the wake region. These
results indicate that the strength of Inequillbrlum turbulence does not
necessarily depend on the turbulence intensity. It can be seen from eq.
(ii) that the rapidly varying _t/ep in the forward region of the Jet will
influence the fluid flow significantly. It takes a while for large eddies
to cascade to smaller eddies. The large ratio of _/k t in the wake region
of the jet is caused by the large eddies convected from the Upstream region
and those generated in the wake region, see Fig. 6-(a).
17
The calculated vertical velocity profiles in the vicinity of the Jet exit
are compared with the measured data in Fig. 7. The slightly higher
w-velocity for x/D>2 is caused by the north boundary which is not located
far enough from the jet exit. Due to the prescribed velocity profile (which
is the same as that of the inlet plane) on the north boundary, the Jet can
not expand freely into the north direction and thus the excess mass flow
rate along the symmetry plane cause the over-predlcted w-veloclty in the
region.
It is shown in Fig. 8 that the calculated tangential velocity profiles
along the x-coordlnate direction on the symmetry plane are in good
agreement with the measured data. The reversed flow behind the Jet indicate
that the development of the tangential velocity along the crossflow
direction is similar to that of the flow over a circular cylinder. However,
the u-veloclty in front of the Jet is not brought to zero due to the
compliance of the Jet.
The u-veloclty profiles at four downstream locations are shown in Fig. 9.
The complexity of the u-veloclty profiles are caused by the deflected Jet
and the separated crossflow behind the Jet. It can be seen in the figure
that the calculated results are in good agreement with the measured data
qualitatively and quantitatively.
The calculated turbulent kinetic energy distribution along the x-axis of
the symmetry plane at z/D-0.75 is compared with the measured data in Fig.
i0. It can be seen in the figure that the trend of the turbulent kinetic
energy distribution is in excellent agreement with the measured data even
though the turbulence intensity is under-predlcted.
The calculated concentration profiles at three downstream locations on
the symmetry plane are shown in Fig. ii, and the concentration contour plot
18
at x/d-8 is shown in Fig. 12-(b). The shape and the peak locations of the
calculated concentration profiles are in very good agreement with the
measured data. The slightly smaller magnitude of the concentration is
caused by the coarse grid inaccuracy in the far downstream region. For
example, the concentration level near (x/D, y/D, z/D)=(8.0, 0.7, 2.1) shown
in Fig. 12-(b) is slightly higher than the measured data, and hence the
concentration level on the symmetry plane is slightly under-predlcted.
The calculated u-veloclty, concentration, and turbulent kinetic energy
contours are shown in Figs. 12-(a), (b), and (c), respectively. It can be
found in Crabb et al. [i] that the present numerical results are in
ex_e_lent agreement with the experimentally obtained contour plots. The
slight difference between the calculated and the measured u-veloclty
contour plots in the vicinity of z/D=4.5 is again attributed to the coarse
grid inaccuracy in the region. Note that the peak concentration occurs in
the region where u-velocity is minimum. This trend indicate that the
turbulent transport of the concentration is significantly different from
that of the momentum. The concentration contour plot exhibits a strong
similarity with the turbulent kinetic energy contour plot. This trend is
also in excellent agreement with the experimentally observed distributions
of the concentration and the turbulent kinetic energy [I]. The excellent
agreement between the calculated and the measured contour plots indicates
that the M-S turbulence model can correctly resolve the turbulent
transports of mass and momentum and that the turbulent transport of mass
and momentum depends strongly on the inequilibrium turbulence. It is not
clear as yet if any single-tlme-scale turbulence model can correctly
resolve the concentration field as the M-S turbulence model can.
The three-dimenslonal particle trajectories are shown in Fig. 13. It can
19
be seen in the figure that the fluid particles passing near the jet exit
are most easily entrained to the jet. It is also shownin the figure that
the fluid particles near the jet edge carry less momentumand hence these
particles are quickly entrained to the helical vortices in the wake region
of the jet. The particle trajectories showthat the large eddy mixing
occurs in the wide region of the jet edge and that the fluid particles in
the center region of the jet does not mix easily with the crossflow. The
concentration profiles shownin Fig. 12 also indicate the sametrend of the
large eddy mixing.
CONCLUSIONS AND DISCUSSION
It has been shown that a strong inequilibrium turbulence field is
characterized by the shape and the frequency domain of the spectral density
that varies widely in space. The influence of the inequilibrium turbulence
on the development of the mean fluid flow (and consequently, on the
development of the turbulence field itself) can be sensed only through
semi-empirical data. Thus the influence of the inequilibrium turbulence on
the mean fluid flow is more difficult to recognize than other turbulence
phenomena such as the anlsotropy of turbulence. The semi-empirical data
show that the eddy viscosity coefficient becomes smaller in the production
dominated region, and becomes larger in the dissipation dominated region.
In the multiple-time-scale turbulence model, the dependence of the eddy
viscosity coefficient on the inequilibrium turbulence is reflected in the
ratio of the dissipation rate (_t) to the energy transfer rate (_p). In the
simplified Split-spectrum case, the measured spectral density curves show
that the ratio of kp/k t is greater for larger eddies, and becomes smaller
as the large eddies cascade down to smaller eddies. The calculated kp/k t
20
using the multlple-tlme-scale turbulence model also shows the same behavior
as that observed in experiments. The accurate numerical results for a wide
class of complex turbulent flows (including the present jet in the
crossflow) obtained using the present multiple-tlme-scale turbulence model
indicate that the turbulent transport of mass, concentration, and momentum
depends strongly on the inequilibrlum turbulence and that the
multiple-tlme-scale turbulence model correctly resolves the inequilibrlum
turbulence phenomena. It is not clear if slngle-tlme-scale turbulence
models can resolve such inequilibrlum turbulence phenomena as yet.
Numerical results for the three-dimenslonal turbulent flow of a circular
jet in a crossflow show that the jet and the crossflow interact very
strongly with each other in the forward region of the jet and that the
interaction creates a strong inequilibrium turbulence field in the forward
region of the jet. The strong interaction between the jet and the crossflow
at the jet exit also influences the fluid flow and the turbulence field in
the upstream region of the Jet. This results suggests that the upstream
region of the circular Jet needs to be included into the computational
domain in order to obtain accurate numerical results or to assess the
predictive capability of a turbulence model.
The calculated velocity, concentration, and turbulence fields are in good
agreement with the measured data. Both the calculated results and the
measured data show that the Jet in crossflow is characterized by highly
complex velocity, concentration, and turbulence fields that are not usually
found in many other turbulent flows. It is discussed in Crabbet al. [i]
that the weak vortex shedding does not influence the mean fluid flow
significantly. The good comparison between the numerical results and the
measured data is also in agreement with such an observation. The calculated
21
tangential velocity, concentration, and turbulent kinetic energy contours
at a downstream location show that the peak concentration occurs where the
tangential velocity becomes local minimum and that the concentration field
exhibits a close resemblance to the turbulence field. These contour plots
are in excellent agreement with the measured contour plots.
22
REFERENCES
i. D. Crabb, D. F. G. Durao and J. H. Whltelaw, "A Round Jet Normal to a
Crossflow," Journa% of Fluid _nglneering, vol. 103, March, 1981, pp.
142-152.
2. Z. A. Khan, "Opposed Jets in Crossflow," Ph.D. thesis, Mechanical
Engineering Dept., Imperial College of Science and Technology, London,
1982.
3. R. W. Claus and S. P. Vanka, "Multigrld Calculations of a jet in Cross
Flow," AIAA Paper 90-0444, Aerospace Sciences Meeting, Reno, Nevada,
January, 1990.
4. S. V. Patankar, D. Basu and S. Alpay, "Prediction of Three-Dimensional
Velocity Field of a Deflected Turbulent jet," Jgu_nal of Fluid
En_D.gineerlng, vol. 99, no. 4, 1977, pp. 758-762.
5. A. J. White, "The Prediction of the Flow and Heat Transfer in the
Vicinity of a Jet in Crossflow," ASME Paper A81-21108, 1980.
6. S. Syed and L. M. Chlappeta, "Finlte-Difference Methods for Reducing
Numerical diffusion in TEACH-Type Calculations," AIAA paper 85-0057,
1985.
7. J. Andreopoulos, "Measurements in a Jet-Pipe Flow Issuing
Perpendicularly into a Cross Stream," Journal of Fluid Englneerin_,
vol. 104, December, 1982, pp. 493-499.
8. S. J. Kline, B. J. Cantwell and G. M. Lilley, The 1980-1981 AFOSR-HTTM
Stanford Conference on Complex Turbulent Flows, vols. 1-3, Stanford
University, Stanford, California, 1981.
9. Kim, S.-W. and Chen, C.-P., "A Multiple-Time-Scale Turbulence model
Based on Variable Partitioning of the Turbulent Kinetic Energy
Spectrum", Numer_c_l Heat Transfer, Part B, Vol. 16, 1989, pp. 193-211.
23
i0. S.-W. Kim, "Numerical Investigation of an Internal Layer in Turbulent
Flow over a Curved Hill," To appear in Numerical Heat Transfer, also
available as NASA TM-102230, 1989.
ii. S.-W. Kim, "Numerical Investigation of Separated Transonic Turbulent
Flows with a Multlple-Time-Scale Turbulence Model," Numerical Heat
Transfer, Part A, vol. 18, 1990, pp. 149-171.
12. S.-W. Kim, "Calculation of Reattaching Shear Layers in Divergent
Channel with a Multlple-Time-Scale Turbulence Model," To appear in AI_
Journal, also available as AIAA Paper 90-0047, 1990.
13. S.-W. Kim, "A Near-Wall Turbulence Model and Its Application to Ful?.
Developed Turbulent Channel and Pipe Flows," Numerical Heat Transfer,
Part B, Vol. 17, 1990., pp. 101-122.
14. S. V. Patankar, Numerical Heat Transfer and Fluid Flow, McGraw-Hill,
New York, 1980.
15. S.-W. Kim, "Calculations of Separated 3-D Flows with a
Pressure-Staggered Navler-Stokes Equations Solver," NASA CR, In print,
1990.
16. S.-W. Kim, "On the Anomaly of Veloclty-Pressure Decoupllng in
Collocated Mesh," NASA TM, In print, 1990.
17. M. Perlc, R. Kessler and G. Scheurerer, "Comparison of Finlte-Volume
Numerical Methods with Staggered and Collocated Grids," _omputers and
_, vol. 16, no. 4, pp. 389-403, 1988.
18. S. Majumdar, "Role of Underrelaxatlon in Momentum Interpolation for
Calculation of Flow with Nonstaggered Grids," Numerical Heat Transfer,
vol. 13, pp. 125-132, 1988.
19. P. S. Klebanoff, "Characteristics of Turbulence in a Boundary Layer
with Zero Pressure Gradient," NACA Report 1247, 1955.
24
20. I. Wygnanskl and H. Fiedler, "Some Measurements in the Self-Preserving
Jet," Journa% o_ Fluid Mechan%cs, vol. 3, part 3, 1969, pp. 577-612.
21. W. Rodl, The Prediction of Free Boundary Layers by Use of a
Two-Equatlon Model of Turbulence," Ph.D. Thesis, University of London,
London, 1972.
22. B. E Launder, "A Generalized Algebraic Stress Transport Hypothesis,"
AIAA Journal, vol. 20, 1982, pp. 436-437.
23. S.-W. Kim and Y.-S. Chen, "A Finite Element Calculation of Turbulent
Boundary Layer Flows with an Algebraic Stress Turbulence Model,"
_omDuter Methods in Applied Mechanics and Engineering, vol. 66, no. I,
January, 1988, pp. 45-63.
24. K. Hanjelic, B. E. Launder and R. Schlestel, "Multlple-Time-Scale
Concepts in Turbulent Shear Flows" in L. J. S. Bradbury, F. Durst, B.
E. Launder, F. W. Schmidt and J. H. Whltelaw, (eds.), Turbulent Shear
Flows, Vol. 2, Sprlnger-Verlag, New York, 1980, pp. 36-49.
25. V. G. Harris, J. A. H. Graham and S. Corrsln, "Further Experiments in
Nearly Homogeneous Turbulent Shear Flow," JQu_n_l 9f Fluid Mechanics,
Vol. 81, 1977, pp. 657-687.
26. S. Tavoularls and U. Karnlk, "Further Experiments on the Evolution of
Turbulent Stresses and Scales in Uniformly Sheared Turbulence," $ournal
o_ Fluid Mechanics, vol. 204, 1989, pp. 457-478.
25
uo
3.0D 1_ _._D
V 'x
Wj, He
Figure 1.--Nomenclature and computational domain for a
circular jet in crossflow.
E
C
- 0 1 2
log10 K [m-1 ]
(a) Spectral density for inequilibrium turbulent flows,
A: maximum shear location in a circular jet [20],
B: center of a circular jet [20], C: free stream region
of a boundary layer flow in zero pressure gradient [1g].
A
I B
KI
K
kp'.fKKl= oEdK, kt =JK= K1EdK
(b) kp/kt for Inequilibrium turbulent flow, A: Pr/t > 1,
B: Pr/_t = I, C: Pr ,=0.0.
Figure 2.--Spectral density.
26
0
3 _ _:_ _ Planejet[21]
I _'_ \ _ Launder[22]
I _ _'1 _ KimandChen[23]
L "_-_ 0 M-S turbulence model [9]
I I
1 2
Pr/ I
Figure 3.--c1_/c_. f (= t/Cp) profiles, A:Pr/E t > 1,
B: Pr/ t = 1, C: Pr ,, 0.0.
(a) Jet velocity, w/Wj.
i o3, :2,, i \
(b) Pressure, p/0.5 pWj 2
(c) Total pressure (p + 0.5 pw2)/0.5 pW 2.
Figure 5.--Contour plots of the flow field at the jet exit.
(a) Top View.
(b) Perspective view.
Figure 4.--Mesh in the vicinity of jet exit,
_---_._--:. :._.._ _._:
, __.._ -:.__
_'_--_. _s-_-_: -
..... _. ___..
(a) Velocity vector.
02/1
2
(b) Pressure, p/O.5 pU=.
(c) Turbulent kinetic energy (k/0.5 U_2).
Figure 6.--Flow and turbulence fields on the symmetry plane.
2?
5Y
9.5-. ,
(d) Pr/Et.
1.25
1-
0.75-
_:"" 0.5-
0.25"
O" -..Z
-0.25
2-'1
1.25--
1"
0.75"
_:'" 0.5-
0 exP't
! I I I I
0 1 2 3 4
5
x/D
(a) z/O - 0.25.
1.25
x/D
(c) z/D = 1.35.
6
Figure 7.--Normal velocity (w) profiles
on the symmetry plane.
(e),_'ep.
1.5, 35
4.5
4.5
_>" 6,5
(0%_,.
Figure 6.--Concluded
1
0.75
_:" O. 5
0.25
0
-0.25
28
2
1.5-
1
=0.5
0-
-0.5
-2
Ca)
exp't
present
1.5
o 1
0.5
_
-0.5
.'_8 i _ _ _ _ 6 -2
x/D
(a) z/D = 0.75. (b) z/D = 1.35.
(b)
I ! I I I I |
-1 0 1 2 3 4 5 6
x/D
Figure 8.mTangential velocity (u) profiles on the symmetry plane.
_3
2
i
0
O_OH exp't 0.751
------- present
0.25-
L,A, , ,:..q. -._, , . o
1 0 1 0 1 0 1 -2
u/U
0
Figure 9.'Tangential velocity (u) profiles
on the symmetry plane, (a) x/D = 0,
(b) x/D = 0.5, (c) x/D = 0.75, (d) x/D =
1.0.
I I
-1 0
..... exp't
present
I I I I I I
1 2 3 4 5 6
x/D
Figure 10.--Turbulent kinetic energy
profile on the symmetry plane at
z/D = 0.75.
7
6
5
_4
3
2
1
0
0 measured data
M-S result
0.0.25 0.0.25 0.0.25
C
Figure 11.--Concentration profiles at
downstream locations on the symmetry
plane, (a) x/D = 4, (b) x/D = 6, (c) x/D =
8.
29
h
[(a)
,._ 1.06
i
i(b)
F.05
y/D y/D
(c)
1.0012
,,r.0025
,_.'/,-.005
-_-"_,'q_/r. 01
20 1 2
y/D
(a) u-velocity. (b) Concentration. (c) Turbulent
kinetic
energy.
Figure 12.--Contour plots of u-velocity, concentration,
and turbulent kinetic energy.
Figure 1&--Three-dimensional particle trajectories of a jet in a crossflow.
3O
I1|
Report Documentation Page
1. Report No. J 2. Government Accession No. 3. Recipient's Catalog No.
NASA TM - 104343
I
4. T'dle and Subtitle
Calculation of a Circular Jet in Crossflow With a
Multiple-Time-Scale Turbulence Model
7. _or(s)
S.-W. Kim and T.J. Benson
9. Performing Organization Name and Address
National Aeronautics and Space Administration
Lewis Research Center
Cleveland, Ohio 44135 - 3191
12. Sponsodng Agen_ Iqame and Addreu
National Aeronautics and Space Administration
Washington, D.C. 20546 - 0001
5, Report Date
July 1991
6. Performing Organization Code
8. Performing Organization Report No.
E -6117
10. Work Unit No.
505 - 62- 52
11. Conlract or Grant No.
13. Type of Report and Period Covered
Technical Memorandum
14. Sponsoring Agency Code
15. Supplementary Notes
S.-W. Kim, University of Texas at Arlington, Department of Aerospace Engineering, Arlington, Texas 76010 and
NASA Resident Research Associate at NASA Lewis Research Center, (work funded by NASA Cooperative
Agreement NCC3-180). T.J. Benson, Lewis Research Center. Responsible person, S.-W. Kim, (216) 433-6682.
16.
Numerical calculation of a three-dimensional turbulent flow of a jet in a crossflow using a multiple-time-scale
turbulence model is presented. The turbulence in the forward region of the jet is in a stronger inequilibrium state
than that in the wake region of the jet, while the turbulence level in the wake region is higher than that in the front
region. The calculated flow and the concentration fields are in very good agreement with the measured data, and it
indicates that the turbulent transport of mass, concentration and momentum is strongly governed by the
inequilibrium turbulence. The capability of the multiple-time-scale turbulence model to resolve the inequilibrium
turbulence field is also discussed.
17. KeyWords (Suggested by Author(s))
Turbulent jets; Cross flow; 'Turbulence;
Turbulent mixing; Cascades
18. Distribution Statement
Unclassified - Unlimited
Subject Category 34
19. Security Clessif. (ot the report) 20. Security Classif. (of this page) 21. No. ot pages ' ' 22. Prioe"
Unclassified Unclassified 32 A03
NAIAFORM11126 OCT86 "For saleby the NationalTechnicalInformationService,Springfield,Virginia 22161

Vous aimerez peut-être aussi